What is a dna adduct

DNA adducts are chemical modifications of DNA. A chemical modification on the DNA strand may hinder DNA replication (like when you get something caught in your jacket zipper), resulting in abnormal replication and mutation.

The formation of the adduct in cells does not mean an individual will develop cancer. The DNA is constantly being damaged. The body has systems in place which can repair damaged DNA or induce the death of cells with high levels of DNA damage. Continuous exposure to the environments and substances that lead to the formation of adducts (e.g. through poor diet and lifestyle) can in the long-term lead to a greater chance of proliferation of cells with damage in genes that make it susceptible to becoming cancers. This is a process that can take decades.

DNA adducts can be detected using a method called, immunohistochemistry. In this method, a tissue of interest is embedded into media on microscope slides. Then, an antibody specific for the DNA adduct is applied. Next, a specific stain is added to the slides. In places where the antibody has bound to tissue containing DNA adducts, the stain will show a red colour. Parts of the tissue that have no DNA adducts will be coloured blue by the stain.

Diet modulates levels of DNA adducts in our body.

References:

  • Le Leu, R.K., Winter, J.M., Christophersen, C.T., Young , G.P., Humphreys, K.J., Hu, Y., Gratz, S.W., Miller, R.B., Topping, D.L., Bird, A.R., Conlon, M.A. “Butyrylated Starch Intake Can Prevent Red Meat-Induced O6-Methyl-2-Deoxyguanosine Adducts in Human Rectal Tissue: a Randomised Clinical Trial.” British Journal of Nutrition, vol. 114, no. 02, 2015, pp. 220–230., doi:10.1017/s0007114515001750.

The term “bulky DNA adducts” comes from early experiments that employed a 32P-DNA postlabeling approach for damage analysis and refers to all the relatively large aromatic and nonpolar carcinogen DNA adducts, such as those induced from polycyclic aromatic hydrocarbons (PAH) and aromatic amines [3–6].

From: Advances in Clinical Chemistry, 2017

1. La DK, Swenberg JA. DNA adducts: Biological markers of exposure and potential applications to risk assessment. Mutat Res. 1996;365:129–46. [PubMed] [Google Scholar]

2. Farmer P. “Biomarkers of exposure and effect for environmental carcinogens, and their applicability to human molecular epidemiological studies”. Public Health Applications of Human Biomonitoring. U. S. EPA. [Last retrieved on 2011 Jun 22]. Available from: www.digplanet.com/wiki/DNA_adduct .

3. den Engelse L, van Benthem J, Scherer E. Immunocytochemical analysis of in vivo DNA modification. Mutat Res. 1990;233:265–87. [PubMed] [Google Scholar]

4. Gorelick NJ, Wogan GN. Fluoranthene-DNA adducts: Identification and quantification by an HPLC-32P-postlabeling method. Carcinogenesis. 1989;10:1567–77. [PubMed] [Google Scholar]

5. Vahakangas K, Haugen A, Harris CC. An applied synchronous fluorescence spectrophotometric assay to study benzo[a] pyrene-diolepoxide-DNA adducts. Carcinogenesis. 1985;6:1109–15. [PubMed] [Google Scholar]

6. van Schooten FJ, van Leeuwen FE, Hillebrand MJ, de Rijke ME, Hart AA, van Veen HG, et al. Determination of benzo[a] pyrene diol epoxide-DNA adducts in white blood cell DNA from coke-oven workers: The impact of smoking. J Natl Cancer Inst. 1990;82:927–33. [PubMed] [Google Scholar]

7. Hemminki K, Grzybowska E, Chorazy M, Twardowska-Saucha K, Sroczynski JW, Putman KL, et al. DNA adducts in human environmentally exposed to aromatic compounds in an industrial area of Poland. Carcinogenesis. 1990;11:1229–31. [PubMed] [Google Scholar]

8. Hemminki K, Randerath K, Reddy MV, Putman KL, Santella RM, Perera FP, et al. Postlabeling and immunoassay analysis of polycyclic aromatic hydrocarbons – Adducts of deoxyribonucleic acid in white blood cells of foundry workers. Scand J Work Environ Health. 1990;16:158–62. [PubMed] [Google Scholar]

9. Phillips DH, Hemminki K, Alhonen A, Hewer A, Grover PL. Monitoring occupational exposure to carcinogens: Detection by 32P-postlabelling of aromatic DNA adducts in white blood cells from iron foundry workers. Mutat Res. 1988;204:531–41. [PubMed] [Google Scholar]

1. Miller EC. Some current perspectives on chemical carcinogenesis in humans and experimental animals: Presidential Address. Cancer Res. 1978;38:1479–1496. [PubMed] [Google Scholar]

2. Scharer OD. Chemistry and biology of DNA repair. Angew Chem Int Ed Engl. 2003;42:2946–2974. [PubMed] [Google Scholar]

3. Sancar A, Lindsey-Boltz LA, Unsal-Kacmaz K, Linn S. Molecular mechanisms of mammalian DNA repair and the DNA damage checkpoints. Annu Rev Biochem. 2004;73:39–85. [PubMed] [Google Scholar]

4. Wood RD, Mitchell M, Sgouros J, Lindahl T. Human DNA repair genes. Science. 2001;291:1284–1289. [PubMed] [Google Scholar]

5. Stratton MR, Campbell PJ, Futreal PA. The cancer genome. Nature. 2009;458:719–724. [PMC free article] [PubMed] [Google Scholar]

6. Loeb LA, Harris CC. Advances in chemical carcinogenesis: a historical review and prospective. Cancer Res. 2008;68:6863–6872. [PMC free article] [PubMed] [Google Scholar]

7. Rosenquist TA, Grollman AP. Mutational signature of aristolochic acid: Clue to the recognition of a global disease. DNA Repair (Amst) 2016;44:205–211. [PubMed] [Google Scholar]

8. IARC. Tobacco smoke and involuntary smoking. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2004;83:1–1438. [PMC free article] [PubMed] [Google Scholar]

9. IARC. Some aromatic amines, organic dyes, and related exposures. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2010;99:1–658. [PMC free article] [PubMed] [Google Scholar]

10. Stiborova M, Arlt VM, Schmeiser HH. DNA Adducts Formed by Aristolochic Acid Are Unique Biomarkers of Exposure and Explain the Initiation Phase of Upper Urothelial Cancer. Int J Mol Sci. 2017:18. [PMC free article] [PubMed] [Google Scholar]

11. Wogan GN, Kensler TW, Groopman JD. Present and future directions of translational research on aflatoxin and hepatocellular carcinoma. A review. Food Addit Contam Part A Chem Anal Control Expo Risk Assess. 2012;29:249–257. [PMC free article] [PubMed] [Google Scholar]

12. Kensler TW, Roebuck BD, Wogan GN, Groopman JD. Aflatoxin: a 50-year odyssey of mechanistic and translational toxicology. Toxicol Sci. 2011;120(Suppl 1):S28–48. [PMC free article] [PubMed] [Google Scholar]

13. Himmelstein MW, Boogaard PJ, Cadet J, et al. Creating context for the use of DNA adduct data in cancer risk assessment: II. Overview of methods of identification and quantitation of DNA damage. Crit Rev Toxicol. 2009;39:679–694. [PubMed] [Google Scholar]

14. Jarabek AM, Pottenger LH, Andrews LS, et al. Creating context for the use of DNA adduct data in cancer risk assessment: I. Data organization. Crit Rev Toxicol. 2009;39:659–678. [PubMed] [Google Scholar]

15. Liu S, Wang Y. Mass spectrometry for the assessment of the occurrence and biological consequences of DNA adducts. Chem Soc Rev. 2015;44:7829–7854. [PMC free article] [PubMed] [Google Scholar]

16. Tretyakova N, Villalta PW, Kotapati S. Mass spectrometry of structurally modified DNA. Chem Rev. 2013;113:2395–2436. [PMC free article] [PubMed] [Google Scholar]

17. Stanley LA. Drug Metabolism. In: Badal S, Delgoda R, editors. Pharmacognosy, Fundamentals Applications and Strategies. 2017. pp. 527–545. [Google Scholar]

18. Guengerich FP. Cytochrome p450 and chemical toxicology. Chem Res Toxicol. 2008;21:70–83. [PubMed] [Google Scholar]

19. Rendic S, Guengerich FP. Contributions of human enzymes in carcinogen metabolism. Chem Res Toxicol. 2012;25:1316–1383. [PMC free article] [PubMed] [Google Scholar]

20. Guengerich FP, Shimada T. Oxidation of toxic and carcinogenic chemicals by human cytochrome P-450 enzymes. Chem Res Toxicol. 1991;4:391–407. [PubMed] [Google Scholar]

21. Cadet J, Wagner JR. DNA base damage by reactive oxygen species, oxidizing agents, and UV radiation. Cold Spring Harb Perspect Biol. 2013:5. [PMC free article] [PubMed] [Google Scholar]

22. Jancova P, Anzenbacher P, Anzenbacherova E. Phase II drug metabolizing enzymes. Biomed Pap Med Fac Univ Palacky Olomouc Czech Repub. 2010;154:103–116. [PubMed] [Google Scholar]

23. Minchin RF, Reeves PT, Teitel CH, et al. N-and O-acetylation of aromatic and heterocyclic amine carcinogens by human monomorphic and polymorphic acetyltransferases expressed in COS-1 cells. Biochem Biophys Res Commun. 1992;185:839–844. [PubMed] [Google Scholar]

24. Thier R, Pemble SE, Kramer H, Taylor JB, Guengerich FP, Ketterer B. Human glutathione S-transferase T1-1 enhances mutagenicity of 1,2-dibromoethane, dibromomethane and 1,2,3,4-diepoxybutane in Salmonella typhimurium. Carcinogenesis. 1996;17:163–166. [PubMed] [Google Scholar]

25. Miller JA, Surh YJ, Liem A, Miller EC. Electrophilic sulfuric acid ester metabolites of hydroxy-methyl aromatic hydrocarbons as precursors of hepatic benzylic DNA adducts in vivo. Adv Exp Med Biol. 1991;283:555–567. [PubMed] [Google Scholar]

26. Regan SL, Maggs JL, Hammond TG, Lambert C, Williams DP, Park BK. Acyl glucuronides: the good, the bad and the ugly. Biopharm Drug Dispos. 2010;31:367–395. [PubMed] [Google Scholar]

27. IARC. Some non-heterocyclic polycyclic aromatic hydrocarbons and some related exposures. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2010;92:1–853. [PMC free article] [PubMed] [Google Scholar]

28. IARC. Diesel and Gasoline Engine Exhausts and Some Nitroarenes. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2014;105:9–699. [PMC free article] [PubMed] [Google Scholar]

29. IARC. Smokeless Tobacco and Some Tobacco-Specific N-Nitrosamines. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2007:89. [PMC free article] [PubMed] [Google Scholar]

30. IARC. Some Traditional Herbal Medicines, Some Mycotoxins, Naphthalene and Styrene. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2002:82. [PMC free article] [PubMed] [Google Scholar]

31. IARC. IARC monographs on the evaluation of the carcinogenic risk of chemicals to man: some aziridines, N-, S- & O-mustards and selenium. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 1975;9:1–268. [PubMed] [Google Scholar]

32. IARC. Overall evaluations of carcinogenicity: an updating of IARC Monographs volumes 1 to 42. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 1987;7:1–440. [PubMed] [Google Scholar]

33. IARC. Some Naturally Occurring Substances: Food Items and Constituents, Heterocyclic Aromatic Amines and Mycotoxins. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 1993;56:1–599. [Google Scholar]

34. Wyatt MD, Pittman DL. Methylating agents and DNA repair responses: Methylated bases and sources of strand breaks. Chem Res Toxicol. 2006;19:1580–1594. [PMC free article] [PubMed] [Google Scholar]

35. Singer B. In vivo formation and persistence of modified nucleosides resulting from alkylating agents. Environ Health Perspect. 1985;62:41–48. [PMC free article] [PubMed] [Google Scholar]

36. Guengerich FP, McCormick WA, Wheeler JB. Analysis of the kinetic mechanism of haloalkane conjugation by mammalian theta-class glutathione transferases. Chem Res Toxicol. 2003;16:1493–1499. [PubMed] [Google Scholar]

37. Wheeler JB, Stourman NV, Armstrong RN, Guengerich FP. Conjugation of haloalkanes by bacterial and mammalian glutathione transferases: mono- and vicinal dihaloethanes. Chem Res Toxicol. 2001;14:1107–1117. [PubMed] [Google Scholar]

38. Turesky RJ, Le Marchand L. Metabolism and biomarkers of heterocyclic aromatic amines in molecular epidemiology studies: lessons learned from aromatic amines. Chem Res Toxicol. 2011;24:1169–1214. [PMC free article] [PubMed] [Google Scholar]

39. Haglund J, Henderson AP, Golding BT, Tornqvist M. Evidence for phosphate adducts in DNA from mice treated with 4-(N-Methyl-N-nitrosamino)-1-(3-pyridyl)-1-butanone (NNK) Chem Res Toxicol. 2002;15:773–779. [PubMed] [Google Scholar]

40. Ma B, Zarth AT, Carlson ES, et al. Identification of More Than One Hundred Structurally Unique DNA-Phosphate Adducts Formed During Rat Lung Carcinogenesis by the Tobacco-Specific Nitrosamine 4-(Methylnitrosamino)-1-(3-pyridyl)-1-butanone. Carcinogenesis. 2017 [PMC free article] [PubMed] [Google Scholar]

41. Gaskell M, Kaur B, Farmer PB, Singh R. Detection of phosphodiester adducts formed by the reaction of benzo[a]pyrene diol epoxide with 2′-deoxynucleotides using collision-induced dissociation electrospray ionization tandem mass spectrometry. Nucleic Acids Res. 2007;35:5014–5027. [PMC free article] [PubMed] [Google Scholar]

42. Dedon PC. The chemical toxicology of 2-deoxyribose oxidation in DNA. Chem Res Toxicol. 2008;21:206–219. [PubMed] [Google Scholar]

43. Lawley PD. Some chemical aspects of dose-response relationships in alkylation mutagenesis. Mutat Res. 1974;23:283–295. [PubMed] [Google Scholar]

44. Pullman B, Pullman A. Nucleophilicity of DNA. Relation to Chemical Carcinogenesis. In: Ts’o POP, Gelboin H, editors. Carcinogenesis: fundamental mechanisms and environmental effects. 1980. pp. 55–66. [Google Scholar]

45. Saffhill R, Margison GP, O’Connor PJ. Mechanisms of carcinogenesis induced by alkylating agents. Biochim Biophys Acta. 1985;823:111–145. [PubMed] [Google Scholar]

46. La DK, Swenberg JA. DNA adducts: biological markers of exposure and potential applications to risk assessment. Mutat Res. 1996;365:129–146. [PubMed] [Google Scholar]

47. Marzilli LA, Wang D, Kobertz WR, Essigmann JM, Vouros P. Mass spectral identification and positional mapping of aflatoxin B1-guanine adducts in oligonucleotides. J Am Soc Mass Spectrom. 1998;9:676–682. [PubMed] [Google Scholar]

48. Chawanthayatham S, Valentine CC, 3rd, Fedeles BI, et al. Mutational spectra of aflatoxin B1 in vivo establish biomarkers of exposure for human hepatocellular carcinoma. P Natl Acad Sci USA. 2017;114:E3101–E3109. [PMC free article] [PubMed] [Google Scholar]

49. Gruppi F, Hejazi L, Christov PP, Krishnamachari S, Turesky RJ, Rizzo CJ. Characterization of nitrogen mustard formamidopyrimidine adduct formation of bis(2-chloroethyl)ethylamine with calf thymus DNA and a human mammary cancer cell line. Chem Res Toxicol. 2015;28:1850–1860. [PMC free article] [PubMed] [Google Scholar]

50. Kohda K, Tada M, Kasai H, Nishimura S, Kawazoe Y. Formation of 8-hydroxyguanine residues in cellular DNA exposed to the carcinogen 4-nitroquinoline 1-oxide. Biochem Biophys Res Commun. 1986;139:626–632. [PubMed] [Google Scholar]

51. Humphreys WG, Kadlubar FF, Guengerich FP. Mechanism of C8 alkylation of guanine residues by activated arylamines: evidence for initial adduct formation at the N7 position. P Natl Acad Sci USA. 1992;89:8278–8282. [PMC free article] [PubMed] [Google Scholar]

52. Beland FAKFF. Metabolic Activation and DNA Adducts of Aromatic Amines and Nitroaromatic Hydrocarbons. In: Cooper CS, Grover PL, editors. Chemical Carcinogenesis and Mutagenesis I. 1. Vol. 94. Springer; Berlin, Heidelberg: 1990. pp. 267–325. [Google Scholar]

53. Turesky RJ, Vouros P. Formation and analysis of heterocyclic aromatic amine-DNA adducts in vitro and in vivo. J Chromatogr B. 2004;802:155–166. [PubMed] [Google Scholar]

54. Kovacic P, Somanathan R. Nitroaromatic compounds: Environmental toxicity, carcinogenicity, mutagenicity, therapy and mechanism. J Appl Toxicol. 2014;34:810–824. [PubMed] [Google Scholar]

55. IARC. IARC monographs on the evaluation of carcinogenic risks to humans. Volume 97. 1,3-butadiene, ethylene oxide and vinyl halides (vinyl fluoride, vinyl chloride and vinyl bromide) IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2008;97:3–471. [PMC free article] [PubMed] [Google Scholar]

56. Pottenger LH, Andrews LS, Bachman AN, et al. An organizational approach for the assessment of DNA adduct data in risk assessment: case studies for aflatoxin B1, tamoxifen and vinyl chloride. Crit Rev Toxicol. 2014;44:348–391. [PubMed] [Google Scholar]

57. Medeiros MH. Exocyclic DNA adducts as biomarkers of lipid oxidation and predictors of disease. Challenges in developing sensitive and specific methods for clinical studies. Chem Res Toxicol. 2009;22:419–425. [PubMed] [Google Scholar]

58. Michaelson-Richie ED, Ming X, Codreanu SG, et al. Mechlorethamine-induced DNA-protein cross-linking in human fibrosarcoma (HT1080) cells. J Proteome Res. 2011;10:2785–2796. [PMC free article] [PubMed] [Google Scholar]

59. Malayappan B, Johnson L, Nie B, et al. Quantitative high-performance liquid chromatography-electrospray ionization tandem mass spectrometry analysis of bis-N7-guanine DNA-DNA cross-links in white blood cells of cancer patients receiving cyclophosphamide therapy. Anal Chem. 2010;82:3650–3658. [PubMed] [Google Scholar]

60. Dedon PC, Tannenbaum SR. Reactive nitrogen species in the chemical biology of inflammation. Arch Biochem Biophys. 2004;423:12–22. [PubMed] [Google Scholar]

61. Steenken S. Purine-Bases, Nucleosides, and Nucleotides - Aqueous-Solution Redox Chemistry and Transformation Reactions of Their Radical Cations and E- and Oh Adducts. Chem Rev. 1989;89:503–520. [Google Scholar]

62. Breen AP, Murphy JA. Reactions of oxyl radicals with DNA. Free Radic Biol Med. 1995;18:1033–1077. [PubMed] [Google Scholar]

63. Kamiya H. Mutagenic potentials of damaged nucleic acids produced by reactive oxygen/nitrogen species: approaches using synthetic oligonucleotides and nucleotides: survey and summary. Nucleic Acids Res. 2003;31:517–531. [PMC free article] [PubMed] [Google Scholar]

64. Cheng Q, Gu J, Compaan KR, Schaefer HF., 3rd Isoguanine formation from adenine. Chemistry. 2012;18:4877–4886. [PubMed] [Google Scholar]

65. Cao H, Wang Y. Quantification of oxidative single-base and intrastrand cross-link lesions in unmethylated and CpG-methylated DNA induced by Fenton-type reagents. Nucleic Acids Res. 2007;35:4833–4844. [PMC free article] [PubMed] [Google Scholar]

66. Brierley DJ, Martin SA. Oxidative stress and the DNA mismatch repair pathway. Antioxid Redox Signal. 2013;18:2420–2428. [PubMed] [Google Scholar]

67. Madugundu GS, Cadet J, Wagner JR. Hydroxyl-radical-induced oxidation of 5-methylcytosine in isolated and cellular DNA. Nucleic Acids Res. 2014;42:7450–7460. [PMC free article] [PubMed] [Google Scholar]

68. Zhang Q, Wang Y. Independent generation of 5-(2′-deoxycytidinyl)methyl radical and the formation of a novel cross-link lesion between 5-methylcytosine and guanine. J Am Chem Soc. 2003;125:12795–12802. [PubMed] [Google Scholar]

69. Yin H, Xu L, Porter NA. Free radical lipid peroxidation: mechanisms and analysis. Chem Rev. 2011;111:5944–5972. [PubMed] [Google Scholar]

70. Tudek B, Zdzalik-Bielecka D, Tudek A, Kosicki K, Fabisiewicz A, Speina E. Lipid peroxidation in face of DNA damage, DNA repair and other cellular processes. Free Radic Biol Med. 2017;107:77–89. [PubMed] [Google Scholar]

71. Blair IA. DNA adducts with lipid peroxidation products. J Biol Chem. 2008;283:15545–15549. [PMC free article] [PubMed] [Google Scholar]

72. Marnett LJ. Oxy radicals, lipid peroxidation and DNA damage. Toxicology. 2002;181–182:219–222. [PubMed] [Google Scholar]

73. Chung FL, Nath RG, Ocando J, Nishikawa A, Zhang L. Deoxyguanosine adducts of t-4-hydroxy-2-nonenal are endogenous DNA lesions in rodents and humans: detection and potential sources. Cancer Res. 2000;60:1507–1511. [PubMed] [Google Scholar]

74. Lee SH, Blair IA. Lipid Peroxide-DNA Adducts. In: Penning T, editor. Chemical Carcinogenesis. New York, NY: Humana Press; 2011. pp. 227–244. [Google Scholar]

75. Delaney JC, Essigmann JM. Biological properties of single chemical-DNA adducts: a twenty year perspective. Chem Res Toxicol. 2008;21:232–252. [PMC free article] [PubMed] [Google Scholar]

76. Shrivastav N, Li D, Essigmann JM. Chemical biology of mutagenesis and DNA repair: cellular responses to DNA alkylation. Carcinogenesis. 2010;31:59–70. [PMC free article] [PubMed] [Google Scholar]

77. Swann PF. Why do O6-alkylguanine and O4-alkylthymine miscode? The relationship between the structure of DNA containing O6-alkylguanine and O4-alkylthymine and the mutagenic properties of these bases. Mutat Res. 1990;233:81–94. [PubMed] [Google Scholar]

78. Preston BD, Singer B, Loeb LA. Mutagenic potential of O4-methylthymine in vivo determined by an enzymatic approach to site-specific mutagenesis. P Natl Acad Sci USA. 1986;83:8501–8505. [PMC free article] [PubMed] [Google Scholar]

79. Singer B. O-alkyl pyrimidines in mutagenesis and carcinogenesis: occurrence and significance. Cancer Res. 1986;46:4879–4885. [PubMed] [Google Scholar]

80. Wang P, Amato NJ, Zhai Q, Wang Y. Cytotoxic and mutagenic properties of O4-alkylthymidine lesions in Escherichia coli cells. Nucleic Acids Res. 2015;43:10795–10803. [PMC free article] [PubMed] [Google Scholar]

81. Rodriguez H, Loechler EL. Mutational specificity of the (+)-anti-diol epoxide of benzo[a]pyrene in a supF gene of an Escherichia coli plasmid: DNA sequence context influences hotspots, mutagenic specificity and the extent of SOS enhancement of mutagenesis. Carcinogenesis. 1993;14:373–383. [PubMed] [Google Scholar]

82. Wei SJ, Chang RL, Bhachech N, et al. Dose-dependent differences in the profile of mutations induced by (+)-7R,8S-dihydroxy-9S,10R-epoxy-7,8,9,10-tetrahydrobenzo(a)pyrene in the coding region of the hypoxanthine (guanine) phosphoribosyltransferase gene in Chinese hamster V-79 cells. Cancer Res. 1993;53:3294–3301. [PubMed] [Google Scholar]

83. Wei SJ, Chang RL, Wong CQ, et al. Dose-dependent differences in the profile of mutations induced by an ultimate carcinogen from benzo[a]pyrene. P Natl Acad Sci USA. 1991;88:11227–11230. [PMC free article] [PubMed] [Google Scholar]

84. Wei SJ, Chang RL, Hennig E, et al. Mutagenic selectivity at the HPRT locus in V-79 cells: comparison of mutations caused by bay-region benzo[a]pyrene 7,8-diol-9,-10-epoxide enantiomers with high and low carcinogenic activity. Carcinogenesis. 1994;15:1729–1735. [PubMed] [Google Scholar]

85. Schiltz M, Cui XX, Lu YP, et al. Characterization of the mutational profile of (+)-7R,8S-dihydroxy-9S, 10R-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene at the hypoxanthine (guanine) phosphoribosyltransferase gene in repair-deficient Chinese hamster V-H1 cells. Carcinogenesis. 1999;20:2279–2286. [PubMed] [Google Scholar]

86. Rodriguez H, Loechler EL. Mutagenesis by the (+)-anti-diol epoxide of benzo[a]pyrene: what controls mutagenic specificity? Biochem. 1993;32:1759–1769. [PubMed] [Google Scholar]

87. Shukla R, Geacintov NE, Loechler EL. The major, N2-dG adduct of (+)-anti-B[a]PDE induces G->A mutations in a 5′-AGA-3′ sequence context. Carcinogenesis. 1999;20:261–268. [PubMed] [Google Scholar]

88. Shukla R, Liu T, Geacintov NE, Loechler EL. The major, N2-dG adduct of (+)-anti-B[a]PDE shows a dramatically different mutagenic specificity (predominantly, G -> A) in a 5′-CGT-3′ sequence context. Biochem. 1997;36:10256–10261. [PubMed] [Google Scholar]

89. Bichara M, Fuchs RP. DNA binding and mutation spectra of the carcinogen N-2-aminofluorene in Escherichia coli. A correlation between the conformation of the premutagenic lesion and the mutation specificity. J Mol Biol. 1985;183:341–351. [PubMed] [Google Scholar]

90. Fuchs RP, Schwartz N, Daune MP. Hot spots of frameshift mutations induced by the ultimate carcinogen N-acetoxy-N-2-acetylaminofluorene. Nature. 1981;294:657–659. [PubMed] [Google Scholar]

91. Patel DJ, Mao B, Gu Z, et al. Nuclear magnetic resonance solution structures of covalent aromatic amine-DNA adducts and their mutagenic relevance. Chem Res Toxicol. 1998;11:391–407. [PubMed] [Google Scholar]

92. Barbin A, Bartsch H, Leconte P, Radman M. Studies on the miscoding properties of 1,N6-ethenoadenine and 3,N4-ethenocytosine, DNA reaction products of vinyl chloride metabolites, during in vitro DNA synthesis. Nucleic Acids Res. 1981;9:375–387. [PMC free article] [PubMed] [Google Scholar]

93. Basu AK, Wood ML, Niedernhofer LJ, Ramos LA, Essigmann JM. Mutagenic and genotoxic effects of three vinyl chloride-induced DNA lesions: 1,N6-ethenoadenine, 3,N4-ethenocytosine, and 4-amino-5-(imidazol-2-yl)imidazole. Biochem. 1993;32:12793–12801. [PubMed] [Google Scholar]

94. Spengler S, Singer B. Transcriptional errors and ambiguity resulting from the presence of 1,N6-ethenoadenosine or 3,N4-ethenocytidine in polyribonucleotides. Nucleic Acids Res. 1981;9:365–373. [PMC free article] [PubMed] [Google Scholar]

95. Singer B, Abbott LG, Spengler SJ. Assessment of mutagenic efficiency of two carcinogen-modified nucleosides, 1,N6-ethenodeoxyadenosine and O4-methyldeoxythymidine, using polymerases of varying fidelity. Carcinogenesis. 1984;5:1165–1171. [PubMed] [Google Scholar]

96. Singer B, Spengler SJ, Chavez F, Kusmierek JT. The vinyl chloride-derived nucleoside, N2,3-ethenoguanosine, is a highly efficient mutagen in transcription. Carcinogenesis. 1987;8:745–747. [PubMed] [Google Scholar]

97. Singer B, Kusmierek JT, Folkman W, Chavez F, Dosanjh MK. Evidence for the mutagenic potential of the vinyl chloride induced adduct, N2, 3-etheno-deoxyguanosine, using a site-directed kinetic assay. Carcinogenesis. 1991;12:745–747. [PubMed] [Google Scholar]

98. Cheng KC, Preston BD, Cahill DS, Dosanjh MK, Singer B, Loeb LA. The vinyl chloride DNA derivative N2,3-ethenoguanine produces G----A transitions in Escherichia coli. P Natl Acad Sci USA. 1991;88:9974–9978. [PMC free article] [PubMed] [Google Scholar]

99. Mroczkowska MM, Kusmierek JT. Miscoding potential of N2,3-ethenoguanine studied in an Escherichia coli DNA-dependent RNA polymerase in vitro system and possible role of this adduct in vinyl chloride-induced mutagenesis. Mutagenesis. 1991;6:385–390. [PubMed] [Google Scholar]

100. Barbin A, Laib RJ, Bartsch H. Lack of miscoding properties of 7-(2-oxoethyl)guanine, the major vinyl chloride-DNA adduct. Cancer Res. 1985;45:2440–2444. [PubMed] [Google Scholar]

101. Chen B, Liu L, Castonguay A, Maronpot RR, Anderson MW, You M. Dose-dependent ras mutation spectra in N-nitrosodiethylamine induced mouse liver tumors and 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone induced mouse lung tumors. Carcinogenesis. 1993;14:1603–1608. [PubMed] [Google Scholar]

102. Kawano R, Takeshima Y, Inai K. Effects of K-ras gene mutations in the development of lung lesions induced by 4-(N-methyl-n-nitrosamino)-1-(3-pyridyl)-1-butanone in A/J mice. Jpn J Cancer Res. 1996;87:44–50. [PMC free article] [PubMed] [Google Scholar]

103. Margison GP, Santibanez Koref MF, Povey AC. Mechanisms of carcinogenicity/chemotherapy by O6-methylguanine. Mutagenesis. 2002;17:483–487. [PubMed] [Google Scholar]

104. Peterson LA, Thomson NM, Crankshaw DL, Donaldson EE, Kenney PJ. Interactions between methylating and pyridyloxobutylating agents in A/J mouse lungs: implications for 4-(methylnitrosamino)-1-(3-pyridyl)-1-butanone-induced lung tumorigenesis. Cancer Res. 2001;61:5757–5763. [PubMed] [Google Scholar]

105. Peterson LA, Hecht SS. O6-Methylguanine Is a Critical Determinant of 4-(Methylnitrosamino)-1-(3-pyridyl)-1-butanone Tumorigenesis in A/J Mouse Lung. Cancer Res. 1991;51:5557–5564. [PubMed] [Google Scholar]

106. Westra WH, Baas IO, Hruban RH, et al. K-ras oncogene activation in atypical alveolar hyperplasias of the human lung. Cancer Res. 1996;56:2224–2228. [PubMed] [Google Scholar]

107. Mills NE, Fishman CL, Rom WN, Dubin N, Jacobson DR. Increased prevalence of K-ras oncogene mutations in lung adenocarcinoma. Cancer Res. 1995;55:1444–1447. [PubMed] [Google Scholar]

108. Rodenhuis S, Slebos RJ. Clinical significance of ras oncogene activation in human lung cancer. Cancer Res. 1992;52:2665s–2669s. [PubMed] [Google Scholar]

109. Noda N, Matsuzoe D, Konno T, Kawahara K, Yamashita Y, Shirakusa T. K-ras gene mutations in non-small cell lung cancer in Japanese. Oncol Rep. 2001;8:889–892. [PubMed] [Google Scholar]

110. Singer B, Essigmann JM. Site-specific mutagenesis: retrospective and prospective. Carcinogenesis. 1991;12:949–955. [PubMed] [Google Scholar]

111. Moriya M. Single-stranded shuttle phagemid for mutagenesis studies in mammalian cells: 8-oxoguanine in DNA induces targeted G.C-->T. A transversions in simian kidney cells. P Natl Acad Sci USA. 1993;90:1122–1126. [PMC free article] [PubMed] [Google Scholar]

112. Moriya M, Zhang W, Johnson F, Grollman AP. Mutagenic potency of exocyclic DNA adducts: marked differences between Escherichia coli and simian kidney cells. P Natl Acad Sci USA. 1994;91:11899–11903. [PMC free article] [PubMed] [Google Scholar]

113. Nesnow S, Ross JA, Stoner GD, Mass MJ. Mechanistic linkage between DNA adducts, mutations in oncogenes and tumorigenesis of carcinogenic environmental polycyclic aromatic hydrocarbons in strain A/J mice. Toxicology. 1995;105:403–413. [PubMed] [Google Scholar]

114. Greenblatt MS, Bennett WP, Hollstein M, Harris CC. Mutations in the p53 tumor suppressor gene: clues to cancer etiology and molecular pathogenesis. Cancer Res. 1994;54:4855–4878. [PubMed] [Google Scholar]

115. Hernandez-Boussard TM, Hainaut P. A specific spectrum of p53 mutations in lung cancer from smokers: review of mutations compiled in the IARC p53 database. Environ Health Perspect. 1998;106:385–391. [PMC free article] [PubMed] [Google Scholar]

116. Hollstein M, Sidransky D, Vogelstein B, Harris CC. p53 mutations in human cancers. Science. 1991;253:49–53. [PubMed] [Google Scholar]

117. Denissenko MF, Pao A, Tang M, Pfeifer GP. Preferential formation of benzo[a]pyrene adducts at lung cancer mutational hotspots in P53. Science. 1996;274:430–432. [PubMed] [Google Scholar]

118. Smith LE, Denissenko MF, Bennett WP, et al. Targeting of lung cancer mutational hotspots by polycyclic aromatic hydrocarbons. J Natl Cancer Inst. 2000;92:803–811. [PubMed] [Google Scholar]

119. Pfeifer GP, Denissenko MF, Olivier M, Tretyakova N, Hecht SS, Hainaut P. Tobacco smoke carcinogens, DNA damage and p53 mutations in smoking-associated cancers. Oncogene. 2002;21:7435–7451. [PubMed] [Google Scholar]

120. Feng Z, Hu W, Hu Y, Tang MS. Acrolein is a major cigarette-related lung cancer agent: Preferential binding at p53 mutational hotspots and inhibition of DNA repair. P Natl Acad Sci USA. 2006;103:15404–15409. [PMC free article] [PubMed] [Google Scholar]

121. Hoang ML, Chen CH, Chen PC, et al. Aristolochic Acid in the Etiology of Renal Cell Carcinoma. Cancer Epidemiol Biomarkers Prev. 2016;25:1600–1608. [PMC free article] [PubMed] [Google Scholar]

122. Hoang ML, Chen CH, Sidorenko VS, et al. Mutational signature of aristolochic acid exposure as revealed by whole-exome sequencing. Sci Transl Med. 2013;5:197ra102. [PMC free article] [PubMed] [Google Scholar]

123. Moriya M, Slade N, Brdar B, et al. TP53 Mutational signature for aristolochic acid: an environmental carcinogen. Int J Cancer. 2011;129:1532–1536. [PubMed] [Google Scholar]

124. Groopman JD, Donahue PR, Zhu JQ, Chen JS, Wogan GN. Aflatoxin metabolism in humans: detection of metabolites and nucleic acid adducts in urine by affinity chromatography. P Natl Acad Sci USA. 1985;82:6492–6496. [PMC free article] [PubMed] [Google Scholar]

125. Groopman JD, Hall AJ, Whittle H, et al. Molecular dosimetry of aflatoxin-N7-guanine in human urine obtained in The Gambia, West Africa. Cancer Epidemiol Biomarkers Prev. 1992;1:221–227. [PubMed] [Google Scholar]

126. Groopman JD, Zhu JQ, Donahue PR, et al. Molecular dosimetry of urinary aflatoxin-DNA adducts in people living in Guangxi Autonomous Region, People’s Republic of China. Cancer Res. 1992;52:45–52. [PubMed] [Google Scholar]

127. Foster PL, Eisenstadt E, Miller JH. Base substitution mutations induced by metabolically activated aflatoxin B1. P Natl Acad Sci USA. 1983;80:2695–2698. [PMC free article] [PubMed] [Google Scholar]

128. Puisieux A, Lim S, Groopman J, Ozturk M. Selective targeting of p53 gene mutational hotspots in human cancers by etiologically defined carcinogens. Cancer Res. 1991;51:6185–6189. [PubMed] [Google Scholar]

129. Aguilar F, Hussain SP, Cerutti P. Aflatoxin B1 induces the transversion of G-->T in codon 249 of the p53 tumor suppressor gene in human hepatocytes. P Natl Acad Sci USA. 1993;90:8586–8590. [PMC free article] [PubMed] [Google Scholar]

130. Denissenko MF, Koudriakova TB, Smith L, O’Connor TR, Riggs AD, Pfeifer GP. The p53 codon 249 mutational hotspot in hepatocellular carcinoma is not related to selective formation or persistence of aflatoxin B1 adducts. Oncogene. 1998;17:3007–3014. [PubMed] [Google Scholar]

131. Smela ME, Currier SS, Bailey EA, Essigmann JM. The chemistry and biology of aflatoxin B(1): from mutational spectrometry to carcinogenesis. Carcinogenesis. 2001;22:535–545. [PubMed] [Google Scholar]

132. Bailey EA, Iyer RS, Stone MP, Harris TM, Essigmann JM. Mutational properties of the primary aflatoxin B1-DNA adduct. P Natl Acad Sci USA. 1996;93:1535–1539. [PMC free article] [PubMed] [Google Scholar]

133. Grollman AP. Aristolochic acid nephropathy: Harbinger of a global iatrogenic disease. Environ Mol Mutagen. 2013;54:1–7. [PubMed] [Google Scholar]

134. Hollstein M, Moriya M, Grollman AP, Olivier M. Analysis of TP53 mutation spectra reveals the fingerprint of the potent environmental carcinogen, aristolochic acid. Mutat Res. 2013;753:41–49. [PMC free article] [PubMed] [Google Scholar]

135. Poon SL, Pang ST, McPherson JR, et al. Genome-wide mutational signatures of aristolochic acid and its application as a screening tool. Sci Transl Med. 2013;5:197ra101. [PubMed] [Google Scholar]

136. Schmeiser HH, Nortier JL, Singh R, et al. Exceptionally long-term persistence of DNA adducts formed by carcinogenic aristolochic acid I in renal tissue from patients with aristolochic acid nephropathy. Int J Cancer. 2014;135:502–507. [PubMed] [Google Scholar]

137. Fernando RC, Schmeiser HH, Scherf HR, Wiessler M. Formation and persistence of specific purine DNA adducts by 32P-postlabelling in target and non-target organs of rats treated with aristolochic acid I. IARC Sci Publ. 1993:167–171. [PubMed] [Google Scholar]

138. Singh R, Farmer PB. Liquid chromatography-electrospray ionization-mass spectrometry: the future of DNA adduct detection. Carcinogenesis. 2006;27:178–196. [PubMed] [Google Scholar]

139. Klaene JJ, Sharma VK, Glick J, Vouros P. The analysis of DNA adducts: the transition from 32P-postlabeling to mass spectrometry. Cancer Lett. 2013;334:10–19. [PMC free article] [PubMed] [Google Scholar]

140. Poirier MC, Santella RM, Weston A. Carcinogen macromolecular adducts and their measurement. Carcinogenesis. 2000;21:353–359. [PubMed] [Google Scholar]

141. Muller R, Adamkiewicz J, Rajewsky MF. Immunological detection and quantification of carcinogen-modified DNA components. IARC Sci Publ. 1982:463–479. [PubMed] [Google Scholar]

142. Phillips DH. DNA adducts as markers of exposure and risk. Mutat Res. 2005;577:284–292. [PubMed] [Google Scholar]

143. Phillips DH. Smoking-related DNA and protein adducts in human tissues. Carcinogenesis. 2002;23:1979–2004. [PubMed] [Google Scholar]

144. Talaska G, Schamer M, Skipper P, et al. Detection of carcinogen-DNA adducts in exfoliated urothelial cells of cigarette smokers: association with smoking, hemoglobin adducts, and urinary mutagenicity. Cancer Epidemiol Biomarkers Prev. 1991;1:61–66. [PubMed] [Google Scholar]

145. Wiencke JK, Kelsey KT, Varkonyi A, et al. Correlation of DNA adducts in blood mononuclear cells with tobacco carcinogen-induced damage in human lung. Cancer Res. 1995;55:4910–4914. [PubMed] [Google Scholar]

146. Hecht SS. Oral Cell DNA Adducts as Potential Biomarkers for Lung Cancer Susceptibility in Cigarette Smokers. Chem Res Toxicol. 2017;30:367–375. [PMC free article] [PubMed] [Google Scholar]

147. Randerath K, Reddy MV, Gupta RC. 32P-labeling test for DNA damage. P Natl Acad Sci USA. 1981;78:6162–6129. [PMC free article] [PubMed] [Google Scholar]

148. Randerath K, Randerath E, Agrawal HP, Gupta RC, Schurda ME, Reddy MV. Postlabeling methods for carcinogen-DNA adduct analysis. Environ Health Perspect. 1985;62:57–65. [PMC free article] [PubMed] [Google Scholar]

149. Shibutani S, Kim SY, Suzuki N. 32P-postlabeling DNA damage assays: PAGE, TLC, and HPLC. Methods Mol Biol. 2006;314:307–321. [PubMed] [Google Scholar]

150. Pfau W, Lecoq S, Hughes NC, et al. Separation of 32P-labelled nucleoside 3′,5′-bisphosphate adducts by HPLC. IARC Sci Publ. 1993:233–242. [PubMed] [Google Scholar]

151. Phillips DH. On the origins and development of the 32P-postlabelling assay for carcinogen-DNA adducts. Cancer Lett. 2013;334:5–9. [PubMed] [Google Scholar]

152. Phillips DH. Polycyclic aromatic hydrocarbons in the diet. Mutat Res. 1999;443:139–147. [PubMed] [Google Scholar]

153. Agudo A, Peluso M, Munnia A, et al. Aromatic DNA adducts and risk of gastrointestinal cancers: a case-cohort study within the EPIC-Spain. Cancer Epidemiol Biomarkers Prev. 2012;21:685–692. [PubMed] [Google Scholar]

154. Gilberson T, Peluso ME, Munia A, et al. Aromatic adducts and lung cancer risk in the European Prospective Investigation into Cancer and Nutrition (EPIC) Spanish cohort. Carcinogenesis. 2014;35:2047–2054. [PubMed] [Google Scholar]

155. Ricceri F, Godschalk RW, Peluso M, et al. Bulky DNA adducts in white blood cells: a pooled analysis of 3,600 subjects. Cancer Epidemiol Biomarkers Prev. 2010;19:3174–3181. [PMC free article] [PubMed] [Google Scholar]

156. Ho V, Peacock S, Massey TE, et al. Gene-diet interactions in exposure to heterocyclic aromatic amines and bulky DNA adduct levels in blood leukocytes. Environ Mol Mutagen. 2015;56:609–620. [PubMed] [Google Scholar]

157. Santella RM. Immunological methods for detection of carcinogen-DNA damage in humans. Cancer Epidemiol Biomarkers Prev. 1999;8:733–739. [PubMed] [Google Scholar]

158. Divi RL, Beland FA, Fu PP, et al. Highly sensitive chemiluminescence immunoassay for benzo[a]pyrene-DNA adducts: validation by comparison with other methods, and use in human biomonitoring. Carcinogenesis. 2002;23:2043–2049. [PubMed] [Google Scholar]

159. Mumford JL, Williams K, Wilcosky TC, Everson RB, Young TL, Santella RM. A sensitive color ELISA for detecting polycyclic aromatic hydrocarbon-DNA adducts in human tissues. Mutat Res. 1996;359:171–177. [PubMed] [Google Scholar]

160. Pratt MM, John K, MacLean AB, Afework S, Phillips DH, Poirier MC. Polycyclic aromatic hydrocarbon (PAH) exposure and DNA adduct semi-quantitation in archived human tissues. Int J Environ Res Public Health. 2011;8:2675–2691. [PMC free article] [PubMed] [Google Scholar]

161. al-Atrash J, Zhang YJ, Lin D, Kadlubar FF, Santella RM. Quantitative immunohistochemical analysis of 4-aminobiphenyl-DNA in cultured cells and mice: comparison to gas chromatography/mass spectroscopy analysis. Chem Res Toxicol. 1995;8:747–752. [PubMed] [Google Scholar]

162. Flamini G, Romano G, Curigliano G, et al. 4-Aminobiphenyl-DNA adducts in laryngeal tissue and smoking habits: an immunohistochemical study. Carcinogenesis. 1998;19:353–357. [PubMed] [Google Scholar]

163. Culp SJ, Roberts DW, Talaska G, et al. Immunochemical, 32P-postlabeling, and GC/MS detection of 4-aminobiphenyl-DNA adducts in human peripheral lung in relation to metabolic activation pathways involving pulmonary N-oxidation, conjugation, and peroxidation. Mutat Res. 1997;378:97–112. [PubMed] [Google Scholar]

164. Beland FA, Doerge DR, Churchwell MI, Poirier MC, Schoket B, Marques MM. Synthesis, characterization, and quantitation of a 4-aminobiphenyl-DNA adduct standard. Chem Res Toxicol. 1999;12:68–77. [PubMed] [Google Scholar]

165. Takahashi S, Tamano S, Hirose M, et al. Immunohistochemical demonstration of carcinogen-DNA adducts in tissues of rats given 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine (PhIP): detection in paraffin-embedded sections and tissue distribution. Cancer Res. 1998;58:4307–4313. [PubMed] [Google Scholar]

166. Zhu J, Chang P, Bondy ML, et al. Detection of 2-amino-1-methyl-6-phenylimidazo[4,5-b]-pyridine-DNA adducts in normal breast tissues and risk of breast cancer. Cancer Epidemiol Biomarkers Prev. 2003;12:830–837. [PubMed] [Google Scholar]

167. Tang D, Liu JJ, Bock CH, et al. Racial differences in clinical and pathological associations with PhIP-DNA adducts in prostate. Int J Cancer. 2007;121:1319–1324. [PMC free article] [PubMed] [Google Scholar]

168. Wohlin P, Zeisig M, Gustafsson JA, Moller L. 32P-HPLC analysis of DNA adducts formed in vitro and in vivo by 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine and 2-amino-3,4,8-trimethyl-3H-imidazo[4,5-f]quinoxaline, utilizing an improved adduct enrichment procedure. Chem Res Toxicol. 1996;9:1050–1056. [PubMed] [Google Scholar]

169. Snyderwine EG, Davis CD, Nouso K, Roller PP, Schut HA. 32P-postlabeling analysis of IQ, MeIQx and PhIP adducts formed in vitro in DNA and polynucleotides and found in vivo in hepatic DNA from IQ-, MeIQx- and PhIP-treated monkeys. Carcinogenesis. 1993;14:1389–1395. [PubMed] [Google Scholar]

170. Zhang Y, Chen SY, Hsu T, Santella RM. Immunohistochemical detection of malondialdehyde-DNA adducts in human oral mucosa cells. Carcinogenesis. 2002;23:207–211. [PubMed] [Google Scholar]

171. Leuratti C, Watson MA, Deag EJ, et al. Detection of malondialdehyde DNA adducts in human colorectal mucosa: relationship with diet and the presence of adenomas. Cancer Epidemiol Biomarkers Prev. 2002;11:267–273. [PubMed] [Google Scholar]

172. Maatouk I, Bouaicha N, Plessis MJ, Perin F. Detection by 32P-postlabelling of 8-oxo-7,8-dihydro-2′-deoxyguanosine in DNA as biomarker of microcystin-LR- and nodularin-induced DNA damage in vitro in primary cultured rat hepatocytes and in vivo in rat liver. Mutat Res. 2004;564:9–20. [PubMed] [Google Scholar]

173. Dizdaroglu M. Quantitative determination of oxidative base damage in DNA by stable isotope-dilution mass spectrometry. FEBS Lett. 1993;315:1–6. [PubMed] [Google Scholar]

174. Brown K. Methods for the detection of DNA adducts. Methods Mol Biol. 2012;817:207–230. [PubMed] [Google Scholar]

175. Guo J, Villalta PW, Turesky RJ. Emerging Mass Spectrometric-Based Technologies Designed to Measure DNA Adducts of Carcinogens in Humans. Environ Mol Mutagen. 2017;58:S28–S29. [Google Scholar]

176. Dizdaroglu M, Coskun E, Jaruga P. Measurement of oxidatively induced DNA damage and its repair, by mass spectrometric techniques. Free Radic Res. 2015;49:525–548. [PubMed] [Google Scholar]

177. Sangaraju D, Goggin M, Walker V, Swenberg J, Tretyakova N. NanoHPLC-nanoESI(+)-MS/MS quantitation of bis-N7-guanine DNA-DNA cross-links in tissues of B6C3F1 mice exposed to subppm levels of 1,3-butadiene. Anal Chem. 2012;84:1732–1739. [PMC free article] [PubMed] [Google Scholar]

178. Guo J, Turesky RJ. Human Biomonitoring of DNA Adducts by Ion Trap Multistage Mass Spectrometry. Curr Protoc Nucleic Acid Chem. 2016;66:7 24 21–27 24 25. [PMC free article] [PubMed] [Google Scholar]

179. McLuckey SA, Van Berkel GJ, Glish GL. Tandem mass spectrometry of small, multiply charged oligonucleotides. J Am Soc Mass Spectrom. 1992;3:60–70. [PubMed] [Google Scholar]

180. Parr GR, Fitzgerald MC, Smith LM. Matrix-Assisted Laser Desorption/Ionization Mass-Spectrometry of Synthetic Oligodeoxyribonucleotides. Rapid Commun Mass Spectrom. 1992;6:369–372. [Google Scholar]

181. Klaene JJ, Flarakos C, Glick J, Barret JT, Zarbl H, Vouros P. Tracking matrix effects in the analysis of DNA adducts of polycyclic aromatic hydrocarbons. J Chromatogr A. 2015 [PMC free article] [PubMed] [Google Scholar]

182. Yin RC, Liu SQ, Zhao C, Lu ML, Tang MS, Wang HL. An Ammonium Bicarbonate-Enhanced Stable Isotope Dilution UHPLC-MS/MS Method for Sensitive and Accurate Quantification of Acrolein-DNA Adducts in Human Leukocytes. Anal Chem. 2013;85:3190–3197. [PubMed] [Google Scholar]

183. Friesen MD, Kaderlik K, Lin D, et al. Analysis of DNA adducts of 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine in rat and human tissues by alkaline hydrolysis and gas chromatography/electron capture mass spectrometry: validation by comparison with 32P-postlabeling. Chem Res Toxicol. 1994;7:733–739. [PubMed] [Google Scholar]

184. Chandhary AK, Nokubo M, Marnett LJ, Blair IA. Analysis of the Malondialdehyde-2′-Deoxyguanosine Adduct in Rat-Liver DNA by Gas-Chromatography Electron-Capture Negative Chemical-Ionization Mass-Spectrometry. Biol Mass Spectrom. 1994;23:457–464. [PubMed] [Google Scholar]

185. Foiles PG, Akerkar SA, Carmella SG, et al. Mass spectrometric analysis of tobacco-specific nitrosamine-DNA adducts in smokers and nonsmokers. Chem Res Toxicol. 1991;4:364–368. [PubMed] [Google Scholar]

186. Lin D, Lay JO, Jr, Bryant MS, et al. Analysis of 4-aminobiphenyl-DNA adducts in human urinary bladder and lung by alkaline hydrolysis and negative ion gas chromatography-mass spectrometry. Environ Health Perspect. 1994;102(Suppl 6):11–16. [PMC free article] [PubMed] [Google Scholar]

187. Giese RW. Detection of DNA adducts by electron capture mass spectrometry. Chem Res Toxicol. 1997;10:255–270. [PubMed] [Google Scholar]

188. Ravanat JL, Turesky RJ, Gremaud E, Trudel LJ, Stadler RH. Determination of 8-Oxoguanine in DNA by Gas-Chromatography Mass-Spectrometry and Hplc-Electrochemical Detection - Overestimation of the Background Level of the Oxidized Base by the Gas-Chromatography Mass-Spectrometry Assay. Chem Res Toxicol. 1995;8:1039–1045. [PubMed] [Google Scholar]

189. Wolf SM, Vouros P. Application of capillary liquid chromatography coupled with tandem mass spectrometric methods to the rapid screening of adducts formed by the reaction of N-acetoxy-N-acetyl-2-aminofluorene with calf thymus DNA. Chem Res Toxicol. 1994;7:82–88. [PubMed] [Google Scholar]

190. Tretyakova N, Goggin M, Sangaraju D, Janis G. Quantitation of DNA adducts by stable isotope dilution mass spectrometry. Chem Res Toxicol. 2012;25:2007–2035. [PMC free article] [PubMed] [Google Scholar]

191. Tarun M, Rusling JF. Quantitative measurement of DNA adducts using neutral hydrolysis and LC-MS. Validation of genotoxicity sensors. Anal Chem. 2005;77:2056–2062. [PubMed] [Google Scholar]

192. Smela ME, Hamm ML, Henderson PT, Harris CM, Harris TM, Essigmann JM. The aflatoxin B(1) formamidopyrimidine adduct plays a major role in causing the types of mutations observed in human hepatocellular carcinoma. P Natl Acad Sci USA. 2002;99:6655–6660. [PMC free article] [PubMed] [Google Scholar]

193. Stone MP, Banerjee S, Brown KL, Egli M. Chemistry and Biology of Aflatoxin-DNA Adducts. Frontiers in Nucleic Acids. 2011;1082:147–166. [Google Scholar]

194. Stepanov I, Muzic J, Le CT, et al. Analysis of 4-Hydroxy-1-(3-pyridyl)-1-butanone (HPB)-Releasing DNA Adducts in Human Exfoliated Oral Mucosa Cells by Liquid Chromatography-Electrospray Ionization-Tandem Mass Spectrometry. Chem Res Toxicol. 2013;26:37–45. [PMC free article] [PubMed] [Google Scholar]

195. Goggin M, Swenberg JA, Walker VE, Tretyakova N. Molecular Dosimetry of 1,2,3,4-Diepoxybutane-Induced DNA-DNA Cross-Links in B6C3F1 Mice and F344 Rats Exposed to 1,3-Butadiene by Inhalation. Cancer Res. 2009;69:2479–2486. [PMC free article] [PubMed] [Google Scholar]

196. Bessette EE, Goodenough AK, Langouet S, et al. Screening for DNA adducts by data-dependent constant neutral loss-triple stage mass spectrometry with a linear quadrupole ion trap mass spectrometer. Anal Chem. 2009;81:809–819. [PMC free article] [PubMed] [Google Scholar]

197. Gu D, Turesky RJ, Tao Y, et al. DNA adducts of 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine and 4-aminobiphenyl are infrequently detected in human mammary tissue by liquid chromatography/tandem mass spectrometry. Carcinogenesis. 2012;33:124–130. [PMC free article] [PubMed] [Google Scholar]

198. Xiao S, Guo J, Yun BH, et al. Biomonitoring DNA Adducts of Cooked Meat Carcinogens in Human Prostate by Nano Liquid Chromatography-High Resolution Tandem Mass Spectrometry: Identification of 2-Amino-1-methyl-6-phenylimidazo[4,5-b]pyridine DNA Adduct. Anal Chem. 2016;88:12508–12515. [PMC free article] [PubMed] [Google Scholar]

199. Yun BH, Rosenquist TA, Sidorenko V, et al. Biomonitoring of Aristolactam-DNA Adducts in Human Tissues Using Ultra-Performance Liquid Chromatography/Ion-Trap Mass Spectrometry. Chem Res Toxicol. 2012;25:1119–1131. [PMC free article] [PubMed] [Google Scholar]

200. Guo J, Yun BH, Upadhyaya P, et al. Multiclass Carcinogenic DNA Adduct Quantification in Formalin-Fixed Paraffin-Embedded Tissues by Ultraperformance Liquid Chromatography-Tandem Mass Spectrometry. Anal Chem. 2016;88:4780–4787. [PMC free article] [PubMed] [Google Scholar]

201. Ma B, Zarth AT, Carlson ES, et al. Methyl DNA Phosphate Adduct Formation in Rats Treated Chronically with 4-(Methylnitrosamino)-1-(3-pyridyl)-1-butanone and Enantiomers of Its Metabolite 4-(Methylnitrosamino)-1-(3-pyridyl)-1-butanol. Chem Res Toxicol. 2017 [PMC free article] [PubMed] [Google Scholar]

202. Fenn JB, Mann M, Meng CK, Wong SF, Whitehouse CM. Electrospray ionization for mass spectrometry of large biomolecules. Science. 1989;246:64–71. [PubMed] [Google Scholar]

203. Juraschek R, Dulcks T, Karas M. Nanoelectrospray - More than just a minimized-flow electrospray ionization source. J Am Soc Mass Spectrom. 1999;10:300–308. [PubMed] [Google Scholar]

204. Shen Y, Zhao R, Berger SJ, Anderson GA, Rodriguez N, Smith RD. High-efficiency nanoscale liquid chromatography coupled on-line with mass spectrometry using nanoelectrospray ionization for proteomics. Anal Chem. 2002;74:4235–4249. [PubMed] [Google Scholar]

205. El-Faramawy A, Siu KW, Thomson BA. Efficiency of nano-electrospray ionization. J Am Soc Mass Spectrom. 2005;16:1702–1707. [PubMed] [Google Scholar]

206. Payne AH, Glish GL. Tandem mass spectrometry in quadrupole ion trap and ion cyclotron resonance mass spectrometers. Methods Enzymol. 2005;402:109–148. [PubMed] [Google Scholar]

207. Schwartz JC, Senko MW, Syka JE. A two-dimensional quadrupole ion trap mass spectrometer. J Am Soc Mass Spectrom. 2002;13:659–669. [PubMed] [Google Scholar]

208. Douglas DJ, Frank AJ, Mao D. Linear ion traps in mass spectrometry. Mass Spectrom Rev. 2005;24:1–29. [PubMed] [Google Scholar]

209. McAlister GC, Phanstiel DH, Brumbaugh J, Westphall MS, Coon JJ. Higher-energy collision-activated dissociation without a dedicated collision cell. Mol Cell Proteomics. 2011;10:O111 009456. [PMC free article] [PubMed] [Google Scholar]

210. Tang Y, Kassie F, Qian X, Ansha B, Turesky RJ. DNA Adduct Formation of 2-Amino-9H-pyrido[2,3-b]indole and 2-Amino-3,4-dimethylimidazo[4,5-f]quinoline in Mouse Liver and Extrahepatic Tissues During a Subchronic Feeding Study. Toxicol Sci. 2013 [PMC free article] [PubMed] [Google Scholar]

211. Le Blanc JCY, Hager JW, Ilisiu AMP, Hunter C, Zhong F, Chu I. Unique scanning capabilities of a new hybrid linear ion trap mass spectrometer (Q TRAP) used for high sensitivity proteomics applications. Proteomics. 2003;3:859–869. [PubMed] [Google Scholar]

212. King R, Fernandez-Metzler C. The use of Qtrap technology in drug metabolism. Curr Drug Metab. 2006;7:541–545. [PubMed] [Google Scholar]

213. Boesl U. Time-of-flight mass spectrometry: Introduction to the basics. Mass Spectrom Rev. 2017;36:86–109. [PubMed] [Google Scholar]

214. Pozo OJ, Van Eenoo P, Deventer K, et al. Comparison between triple quadrupole, time of flight and hybrid quadrupole time of flight analysers coupled to liquid chromatography for the detection of anabolic steroids in doping control analysis. Anal Chim Acta. 2011;684:98–111. [PubMed] [Google Scholar]

215. Grimalt S, Sancho JV, Pozo OJ, Hernandez F. Quantification, confirmation and screening capability of UHPLC coupled to triple quadrupole and hybrid quadrupole time-of-flight mass spectrometry in pesticide residue analysis. J Mass Spectrom. 2010;45:421–436. [PubMed] [Google Scholar]

216. Dillen L, Cools W, Vereyken L, et al. Comparison of triple quadrupole and high-resolution TOF-MS for quantification of peptides. Bioanalysis. 2012;4:565–579. [PubMed] [Google Scholar]

217. Wang JJ, Marshall WD, Frazer DG, Law B, Lewis DM. Characterization of DNA adducts from lung tissue of asphalt fume-exposed mice by nanoflow liquid chromatography quadrupole time-of-flight mass spectrometry. Anal Biochem. 2003;322:79–88. [PubMed] [Google Scholar]

218. Gao H, Guo F, Feng F, Yin J, Song M, Wang H. Improved preparation and identification of aristolochic acid-DNA adducts by solid-phase extraction with liquid chromatography-tandem mass spectrometry. J Environ Sci (China) 2009;21:1769–1776. [PubMed] [Google Scholar]

219. Dementeva N, Kokova D, Ponamoreva A, Cherdyntseva N, Kzhyshkowska J. Application of UPLC-ESI-Q-TOF Analysis for Screening of the Carcinogen-Modified DNA-Adducts in the Circulation DNA of Patients with Lung Cancer. Prospects of Fundamental Sciences Development (Pfsd-2016) 2016:1772. [Google Scholar]

220. Ma B, Ruszczak C, Jain V, et al. Optimized Liquid Chromatography Nanoelectrospray-High-Resolution Tandem Mass Spectrometry Method for the Analysis of 4-Hydroxy-1-(3-pyridyl)-1-butanone-Releasing DNA Adducts in Human Oral Cells. Chem Res Toxicol. 2016;29:1849–1856. [PMC free article] [PubMed] [Google Scholar]

221. Gabelica V, Marklund E. Fundamentals of ion mobility spectrometry. Curr Opin Chem Biol. 2018;42:51–59. [PubMed] [Google Scholar]

222. Kafle A, Klaene J, Hall AB, Glick J, Coy SL, Vouros P. A differential mobility spectrometry/mass spectrometry platform for the rapid detection and quantitation of DNA adduct dG-ABP. Rapid Commun Mass Spectrom. 2013;27:1473–1480. [PMC free article] [PubMed] [Google Scholar]

223. Escher BI, Hackermuller J, Polte T, et al. From the exposome to mechanistic understanding of chemical-induced adverse effects. Environ Int. 2017;99:97–106. [PMC free article] [PubMed] [Google Scholar]

224. Buck Louis GM, Smarr MM, Patel CJ. The Exposome Research Paradigm: an Opportunity to Understand the Environmental Basis for Human Health and Disease. Curr Environ Health Rep. 2017;4:89–98. [PMC free article] [PubMed] [Google Scholar]

225. Balbo S, Turesky RJ, Villalta PW. DNA adductomics. Chem Res Toxicol. 2014;27:356–366. [PMC free article] [PubMed] [Google Scholar]

226. Villalta PW, Balbo S. The Future of DNA Adductomic Analysis. Int J Mol Sci. 2017:18. [Google Scholar]

227. Wang P, Fisher D, Rao A, Giese RW. Nontargeted nucleotide analysis based on benzoylhistamine labeling-MALDI-TOF/TOF-MS: discovery of putative 6-oxo-thymine in DNA. Anal Chem. 2012;84:3811–3819. [PMC free article] [PubMed] [Google Scholar]

228. Hemeryck LY, Decloedt AI, Vanden Bussche J, Geboes KP, Vanhaecke L. High resolution mass spectrometry based profiling of diet-related deoxyribonucleic acid adducts. Anal Chim Acta. 2015;892:123–131. [PubMed] [Google Scholar]

229. Chaereboudt J, Esmans EL, Van den Eeckhout EG, Claeys M. Fast Atom Bombardment and Tandem Mass Spectrometry for the Identification of Nucleoside Adducts with Phenyl Glycidyl Ether. Nucleosides & Nucleotides. 1990;9:333–344. [Google Scholar]

230. Kanaly RA, Matsui S, Hanaoka T, Matsuda T. Application of the adductome approach to assess intertissue DNA damage variations in human lung and esophagus. Mutat Res. 2007;625:83–93. [PubMed] [Google Scholar]

231. Chou PH, Kageyama S, Matsuda S, et al. Detection of lipid peroxidation-induced DNA adducts caused by 4-oxo-2(E)-nonenal and 4-oxo-2(E)-hexenal in human autopsy tissues. Chem Res Toxicol. 2010;23:1442–1448. [PubMed] [Google Scholar]

232. Matsuda T, Tao H, Goto M, et al. Lipid peroxidation-induced DNA adducts in human gastric mucosa. Carcinogenesis. 2013;34:121–127. [PubMed] [Google Scholar]

233. Balbo S, Hecht SS, Upadhyaya P, Villalta PW. Application of a high-resolution mass-spectrometry-based DNA adductomics approach for identification of DNA adducts in complex mixtures. Anal Chem. 2014;86:1744–1752. [PMC free article] [PubMed] [Google Scholar]

234. Stornetta A, Villalta PW, Hecht SS, Sturla SJ, Balbo S. Screening for DNA Alkylation Mono and Cross-Linked Adducts with a Comprehensive LC-MS(3) Adductomic Approach. Anal Chem. 2015;87:11706–11713. [PMC free article] [PubMed] [Google Scholar]

235. Gavina JM, Yao C, Feng YL. Recent developments in DNA adduct analysis by mass spectrometry: a tool for exposure biomonitoring and identification of hazard for environmental pollutants. Talanta. 2014;130:475–494. [PubMed] [Google Scholar]

236. Yao C, Feng YL. A nontargeted screening method for covalent DNA adducts and DNA modification selectivity using liquid chromatography-tandem mass spectrometry. Talanta. 2016;159:93–102. [PubMed] [Google Scholar]

237. Van den Driessche B, Van Dongen W, Lemiere F, Esmans EL. Implementation of data-dependent acquisitions in the study of melphalan DNA adducts by miniaturized liquid chromatography coupled to electrospray tandem mass spectrometry. Rapid Commun Mass Spectrom. 2004;18:2001–2007. [PubMed] [Google Scholar]

238. Ishino K, Kato T, Kato M, et al. Comprehensive DNA adduct analysis reveals pulmonary inflammatory response contributes to genotoxic action of magnetite nanoparticles. Int J Mol Sci. 2015;16:3474–3492. [PMC free article] [PubMed] [Google Scholar]

239. Gillet LC, Navarro P, Tate S, et al. Targeted data extraction of the MS/MS spectra generated by data-independent acquisition: a new concept for consistent and accurate proteome analysis. Mol Cell Proteomics. 2012;11:O111 016717. [PMC free article] [PubMed] [Google Scholar]

240. Guo J, Villalta PW, Turesky RJ. Data-Independent Mass Spectrometry Approach for Screening and. Identification of DNA Adducts. Anal Chem. 2017;89:11728–11736. [PMC free article] [PubMed] [Google Scholar]

241. Yun BH, Rosenquist TA, Nikolic J, et al. Human Formalin-Fixed Paraffin-Embedded Tissues: An Untapped Specimen for Biomonitoring of Carcinogen DNA Adducts by Mass Spectrometry. Anal Chem. 2013;85:4251–4258. [PMC free article] [PubMed] [Google Scholar]

242. Cui L, Takahashi S, Tada M, et al. Immunohistochemical Detection of Carcinogen-DNA Adducts in Normal Human Prostate Tissues Transplanted into the Subcutis of Athymic Nude Mice: Results with 2-Amino-1-methyl-6-phenylimidazo[4,5-b]pyridine (PhIP) and 3,2′-Dimethyl-4-aminobiphenyl (DMAB) and. Jpn J Cancer Res. 2000;91:52–58. [PMC free article] [PubMed] [Google Scholar]

243. Zhang Y-J, Weksler BB, Wang L, Schwartz J, Santella RM. Immunohistochemical detection of polycyclic aromatic hydrocarbon-DNA damage in human blood vessels of smokers and non-smokers. Atherosclerosis. 1998;140:325–331. [PubMed] [Google Scholar]

244. Zhang YJ, Geacintov NE, Santella RM. Development of a monoclonal antibody recognizing benzo[c]phenanthrenediol epoxide-DNA adducts: Application to immunohistochemical detection of DNA damage. Chem Res Toxicol. 1997;10:948–952. [PubMed] [Google Scholar]

245. Curigliano G, Zhang Y-J, Wang L-Y, et al. Immunohistochemical quantitation of 4-aminobiphenyl-DNA adducts and p53 nuclear overexpression in T1 bladder cancer of smokers and nonsmokers. Carcinogenesis. 1996;17:911–916. [PubMed] [Google Scholar]

246. Huitfeldt HS, Spangler EF, Hunt JM, Poirier MC. Immunohistochemical localization of DNA adducts in rat liver tissue and phenotypically altered foci during oral administration of 2-acetylaminofluorene. Carcinogenesis. 1986;7:123–129. [PubMed] [Google Scholar]

247. Yun BH, Xiao S, Yao L, et al. A Rapid Throughput Method to Extract DNA from Formalin-Fixed Paraffin-Embedded Tissues for Biomonitoring Carcinogenic DNA Adducts in Humans. Chem Res Toxicol. 2017 [PMC free article] [PubMed] [Google Scholar]

248. Fox CH, Johnson FB, Whiting J, Roller PP. Formaldehyde fixation. J Histochem Cytochem. 1985;33:845–853. [PubMed] [Google Scholar]

249. Kennedy-Darling J, Smith LM. Measuring the Formaldehyde Protein–DNA Cross-Link Reversal Rate. Anal Chem. 2014;86:5678–5681. [PMC free article] [PubMed] [Google Scholar]

250. Fraenkel-Conrat H, Olcott HS. The Reaction of Formaldehyde with Proteins. V. Cross-linking between Amino and Primary Amide or Guanidyl Groups. Journal of the American Chemical Society. 1948;70:2673–2684. [PubMed] [Google Scholar]

251. Hoffman EA, Frey BL, Smith LM, Auble DT. Formaldehyde crosslinking: a tool for the study of chromatin complexes. J Biol Chem. 2015;290:26404–26411. [PMC free article] [PubMed] [Google Scholar]

252. Jackson V. Studies on histone organization in the nucleosome using formaldehyde as a reversible cross-linking agent. Cell. 1978;15:945–954. [PubMed] [Google Scholar]

253. Yun BH, Yao L, Jelakovic B, et al. Formalin-fixed paraffin-embedded tissue as a source for quantitation of carcinogen DNA adducts: aristolochic acid as a prototype carcinogen. Carcinogenesis. 2014;35:2055–2061. [PMC free article] [PubMed] [Google Scholar]

254. IARC. Pharmaceuticals. Volume 100 A. A review of human carcinogens. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 2012;100:1–401. [PMC free article] [PubMed] [Google Scholar]

255. Mottier P, Parisod V, Turesky RJ. Quantitative determination of polycyclic aromatic hydrocarbons in barbecued meat sausages by gas chromatography coupled to mass spectrometry. J Agric Food Chem. 2000;48:1160–1166. [PubMed] [Google Scholar]

256. International Agency for Research on Cancer. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans: Tobacco smoking. Lyon, France: International Agency for Research on Cancer; 1986. [Google Scholar]

257. IARC. Tobacco smoking. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans. 1986:38. [PubMed] [Google Scholar]

258. Neumann HG. Health risk of combustion products: toxicological considerations. Chemosphere. 2001;42:473–479. [PubMed] [Google Scholar]

259. Turesky RJ, Freeman JP, Holland RD, et al. Identification of aminobiphenyl derivatives in commercial hair dyes. Chem Res Toxicol. 2003;16:1162–1173. [PubMed] [Google Scholar]

260. Felton JS, Jagerstad IM, knize MG, Skog K, Wakabayashi K. In Food Borne Carcinogens; Heterocylclic Amines. John Wiley & Sons Ltd; Chichester, England: 2000. [Google Scholar]

261. Sugimura T, Wakabayashi K, Nakagama H, Nagao M. Heterocyclic amines: Mutagens/carcinogens produced during cooking of meat and fish. Cancer Sci. 2004;95:290–299. [PubMed] [Google Scholar]

262. Hecht SS. Progress and challenges in selected areas of tobacco carcinogenesis. Chem Res Toxicol. 2008;21:160–171. [PMC free article] [PubMed] [Google Scholar]

263. Tang D, Kryvenko ON, Wang Y, et al. 2-amino-1-methyl-6-phenylimidazo[4,5-b]pyridine (PhIP)-DNA adducts in benign prostate and subsequent risk for prostate cancer. Int J Cancer. 2013 [PMC free article] [PubMed] [Google Scholar]

264. Tang D, Liu JJ, Rundle A, et al. Grilled meat consumption and PhIP-DNA adducts in prostate carcinogenesis. Cancer Epidemiol Biomarkers Prev. 2007;16:803–808. [PMC free article] [PubMed] [Google Scholar]

265. Cairns J. Mutation selection and the natural history of cancer. Nature. 1975;255:197–200. [PubMed] [Google Scholar]

266. Rojas E, Valverde M, Sordo M, Ostrosky-Wegman P. DNA damage in exfoliated buccal cells of smokers assessed by the single cell gel electrophoresis assay. Mutat Res. 1996;370:115–120. [PubMed] [Google Scholar]

267. Talaska G, Schamer M, Skipper P, et al. Carcinogen-DNA adducts in exfoliated urothelial cells: techniques for noninvasive human monitoring. Environ Health Perspect. 1993;99:289–291. [PMC free article] [PubMed] [Google Scholar]

268. Bessette EE, Spivack SD, Goodenough AK, et al. Identification of carcinogen DNA adducts in human saliva by linear quadrupole ion trap/multistage tandem mass spectrometry. Chem Res Toxicol. 2010;23:1234–1244. [PMC free article] [PubMed] [Google Scholar]

269. Kaderlik KR, Talaska G, DeBord DG, Osorio AM, Kadlubar FF. 4,4′-Methylene-bis(2-chloroaniline)-DNA adduct analysis in human exfoliated urothelial cells by 32P-postlabeling. Cancer Epidemiol Biomarkers Prev. 1993;2:63–69. [PubMed] [Google Scholar]

270. Talaska G, al Juburi AZ, Kadlubar FF. Smoking related carcinogen-DNA adducts in biopsy samples of human urinary bladder: identification of N-(deoxyguanosin-8-yl)-4-aminobiphenyl as a major adduct. P Natl Acad Sci USA. 1991;88:5350–5354. [PMC free article] [PubMed] [Google Scholar]

271. Peters S, Talaska G, Jonsson BA, Kromhout H, Vermeulen R. Polycyclic aromatic hydrocarbon exposure, urinary mutagenicity, and DNA adducts in rubber manufacturing workers. Cancer Epidemiol Biomarkers Prev. 2008;17:1452–1459. [PubMed] [Google Scholar]

272. Vermeulen R, Talaska G, Schumann B, Bos RP, Rothman N, Kromhout H. Urothelial cell DNA adducts in rubber workers. Environ Mol Mutagen. 2002;39:306–313. [PubMed] [Google Scholar]

273. Hsu TM, Zhang YJ, Santella RM. Immunoperoxidase quantitation of 4-aminobiphenyl- and polycyclic aromatic hydrocarbon-DNA adducts in exfoliated oral and urothelial cells of smokers and nonsmokers. Cancer Epidemiol Biomarkers Prev. 1997;6:193–199. [PubMed] [Google Scholar]

274. Fernando RC, Schmeiser HH, Nicklas W, Wiessler M. Detection and quantitation of dG-AAI and dA-AAI adducts by 32P-postlabeling methods in urothelium and exfoliated cells in urine of rats treated with aristolochic acid I. Carcinogenesis. 1992;13:1835–1839. [PubMed] [Google Scholar]

275. Guo L, Wu H, Yue H, Lin S, Lai Y, Cai Z. A novel and specific method for the determination of aristolochic acid-derived DNA adducts in exfoliated urothelial cells by using ultra performance liquid chromatography-triple quadrupole mass spectrometry. J Chromatogr B. 2011;879:153–158. [PubMed] [Google Scholar]

276. Tao Su XL, Yang L, Yun BH, Turesky JR, Grollman A, Dickman K. A noninvasive biomarker for aristolochic acid exposure detected in exfoliated urinary cells. J Am Soc Nephrol. 2013;24:824A. [Google Scholar]

277. Yun BH, Sidorenko VS, Rosenquist TA, Dickman KG, Grollman AP, Turesky RJ. New Approaches for Biomonitoring Exposure to the Human Carcinogen Aristolochic Acid. Toxicol Res (Camb) 2015;4:763–776. [PMC free article] [PubMed] [Google Scholar]

278. Romano G, Sgambato A, Boninsegna A, et al. Evaluation of polycyclic aromatic hydrocarbon-DNA adducts in exfoliated oral cells by an immunohistochemical assay. Cancer Epidemiol Biomarkers Prev. 1999;8:91–96. [PubMed] [Google Scholar]

279. Besaratinia A, Van Straaten HW, Godschalk RW, et al. Immunoperoxidase detection of polycyclic aromatic hydrocarbon-DNA adducts in mouth floor and buccal mucosa cells of smokers and nonsmokers. Environ Mol Mutagen. 2000;36:127–133. [PubMed] [Google Scholar]

280. Balbo S, Meng L, Bliss RL, Jensen JA, Hatsukami DK, Hecht SS. Time course of DNA adduct formation in peripheral blood granulocytes and lymphocytes after drinking alcohol. Mutagenesis. 2012;27:485–490. [PMC free article] [PubMed] [Google Scholar]

281. Chen HJ, Lin WP. Quantitative analysis of multiple exocyclic DNA adducts in human salivary DNA by stable isotope dilution nanoflow liquid chromatography-nanospray ionization tandem mass spectrometry. Anal Chem. 2011;83:8543–8551. [PubMed] [Google Scholar]

282. Law KP, Lim YP. Recent advances in mass spectrometry: data independent analysis and hyper reaction monitoring. Expert Rev Proteomics. 2013;10:551–566. [PubMed] [Google Scholar]

283. Bhattacharya S, Barbacci DC, Shen M, Liu JN, Casale GP. Extraction and purification of depurinated benzo[a]pyrene-adducted DNA bases from human urine by immunoaffinity chromatography coupled with HPLC and analysis by LC/quadrupole ion-trap MS. Chem Res Toxicol. 2003;16:479–486. [PubMed] [Google Scholar]

284. Zhang Q, Aft RL, Gross ML. Estrogen carcinogenesis: specific identification of estrogen-modified nucleobase in breast tissue from women. Chem Res Toxicol. 2008;21:1509–1513. [PMC free article] [PubMed] [Google Scholar]

285. Zayas B, Stillwell SW, Wishnok JS, et al. Detection and quantification of 4-ABP adducts in DNA from bladder cancer patients. Carcinogenesis. 2007;28:342–349. [PubMed] [Google Scholar]

286. Villalta PW, Hochalter JB, Hecht SS. Ultrasensitive High-Resolution Mass Spectrometric Analysis of a DNA Adduct of the Carcinogen Benzo[a]pyrene in Human Lung. Anal Chem. 2017;89:12735–12742. [PMC free article] [PubMed] [Google Scholar]

287. Ma B, Villalta PW, Balbo S, Stepanov I. Analysis of a malondialdehyde-deoxyguanosine adduct in human leukocyte DNA by liquid chromatography nanoelectrospray-high-resolution tandem mass spectrometry. Chem Res Toxicol. 2014;27:1829–1836. [PMC free article] [PubMed] [Google Scholar]


Page 2

Strengths and limitations of the methods for DNA adduct analysis: 32P-postlabeling, immuno-based assay, GC/MS and LC/MS.

AdvantagesLimitations
32P-Postlabeling
  • Requires low μg levels of DNA

  • Detects wide-range of adducts simultaneously

  • Highly sensitive (1 adduct per 1010 nucleotides)

  • Enzymatic digestion and labeling efficiencies vary among adducts

  • Utilizes high levels of radioactive 32P isotope

  • Very labor intensive

  • DNA adduct levels cannot be quantified and structures not identified

Immunoassay/IHC
  • Semi-quantitation achieved through use of calibration curve

  • (IHC) Allows visualization of adduct locations in human tissues or cell types, and can be applied to large numbers of samples

  • Antiserum may cross-react with adducts of similar structure, or cellular components (depending on the epitope)

  • Only semi-quantitative

  • DNA adduct structures are not known

GC/MS or LC/MS
  • Requires low μg levels of DNA

  • Adduct structures are known

  • Absolute quantitation possible (targeted approach)

  • Detects wide-range of adducts simultaneously (untargeted approach)

  • Highly skilled personnel are required

  • Instruments are costly